You are on page 1of 36

RTO-AG-300-V26 3 - 1

Chapter 3 LASER SYSTEMS PERFORMANCE


3.1 GENERAL
A fundamental problem in laser systems performance analysis is determination of the total optical power
that is present at the receiver aperture (case of LADAR and LRF) or LGW seeker (case of LTD) and,
consequently, the total optical power incident on the photosensitive element of the receiver: the detector.
The laser range equation is used to determine the power received under specific conditions and against a
particular target. For laser systems performance analysis specific models are also needed for atmospheric
propagation, target reflection, detection performance, etc.
In general, a laser beam is attenuated as it propagates through the atmosphere. In addition, the beam is often
broadened, defocused, and may even be deflected from its initial propagation direction. The attenuation and
amount of beam alteration depend on the wavelength of operation, output power and characteristics of the
atmosphere. When the output power is low, the effects are linear in behaviour (absorption, scattering, and
atmospheric turbulence are examples of linear effects). On the other hand, when the power is sufficiently
high, new effects are observed that are characterised by non-linear relationships (e.g., thermal blooming,
kinetic cooling, bleaching, and atmospheric breakdown). In both cases, the atmospheric effects can be
significant and severely limit the usefulness of the beam.
Another key element of laser systems performance analysis is the knowledge of target reflection properties.
In general, the reflectivity of a surface can be expressed by two components: the specular component and the
diffuse component. The specular component is the energy that reflects away from the surface at the opposite
of the angle of incidence with the exit beam remaining narrow. The diffuse (Lambertian) component, on the
other hand, is the energy reflected in all directions with a maximum along the normal to the target surface
and falling off as a function of the cosine of the angle off of surface normal.
In most practical cases, target surfaces are very rough at laser wavelengths and, consequently, the diffuse
scattering component frequently dominates (in some cases, however, significant specular components are
observed). Furthermore, most targets exhibit a marked dependency of the overall scattering characteristics
on the illumination incidence angle.
In this chapter, some theoretical background is given of laser systems performance analysis, including
discussions about mission performance requirements, atmospheric propagation and target reflection
properties.
3.2 LASER RANGE EQUATION
The classical forms of the laser range equation, applicable to extended, point and linear (wire type)
targets are presented in Annex B. Furthermore, various considerations are presented relative to laser radar
systems detection performances. Particularly, the signal-to-noise ratio (SNR) equations applicable to both
coherent and incoherent detection laser radar system are presented, and the influence of both background
and system/detector noise terms on the overall systems performance are investigated.
The range equations presented in Annex B assume that the transmitter and receiver are collocated and
have the same optics diameter. In some cases (e.g., for LTD/LGW combinations), these assumptions are
not valid and other forms of the range equation need to be developed.
3.2.1 Range Equation for Airborne LTD/LRF Systems
With reference to the geometry of a typical ground attack mission with laser guided weapons shown in
Figure 3-1, the range performance of an LTD can be estimated using the procedure described below [2].
LASER SYSTEMS PERFORMANCE
3 - 2 RTO-AG-300-V26



Figure 3-1: LTD/LGW Mission Geometry (Vertical Profile).
3.2.1.1 Energy Density on the Target
The laser beam area at a distance R
T
is given by:

( )
4
2
T T L
b
R D
A
+
= (3.1)
where:
D
L
= Transmitted beam diameter (m); and

T
= Output laser beamwidth (rad).
The energy density at the target location (J/m
2
) as a function of transmitted energy (U) is given by:

) (
T Ht w
R
b
e
A
U
F

= (3.2)
This energy density is measured normally to the transmitter line of sight. Using Eq. (3.1), Eq. (3.2) can be
written in the form:

( )
( )
T Ht w
R
T T L
e
R D
U
F



+
=
2
4
(3.3)
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 3


The parameters appearing in the exponential factor are defined as follows:

w
= sea level atmospheric attenuation coefficient; and

Ht
= fractional decrease in w for a path from altitude Ht to sea level.
3.2.1.2 Target Irradiance
The energy (G) of a laser spot that will irradiate a given target surface (A) is that portion passing through
the projected area (A
N
) in the plane orthogonal to the sight line. Therefore, the irradiance of the target
surface can be calculated using the equation:

A
A
F G
N
= (3.4)
and, using the Eq. (3.2):

) (
T HT w
R
b
N
e
A A
UA
G

= (3.5)
or:

) (
2
) (
4
T HT w
R
T T L
N
e
R D A
UA
G



+
= (3.6)
As
t N
A A cos = , we also have:

( )
) (
2
4
cos
T HT w
R
T T L
t
e
R D
U
G


+
= (3.7)
where
t
is the incidence angle to the target surface as measured from the sight line to the target normal.
3.2.1.3 Target Brightness
The brightness of the irradiated target is determined by the irradiance level and by the reflectance
characteristics of the target surface.
The laser energy reaching the target is partially absorbed and partially reflected, either specularly and
diffusely. The probabilities of each of these occurrences are called the coefficients of absorption, specular
reflection, and diffuse reflection, and must satisfy: 1 = + +
d s a
C C C . More details about target reflection
properties are given in successive sections of this chapter. Assuming now that the target is a perfectly
diffuse reflector, with a Lambertian radiation pattern, the brightness (B) is given by:

G
B
T
= (3.8)
where
T
is the target reflectivity.
3.2.1.4 Energy at the Receiver
The energy (E
R
) collected by a receiving aperture observing this target is obtained from the expression:

) (
2
R HR w
R
R
M R
R
e
R
A BA
E

= (3.9)
LASER SYSTEMS PERFORMANCE
3 - 4 RTO-AG-300-V26


where:
A
R
= receiver aperture area;
A
R
/R
R
2

= solid angle subtended by the receiving aperture;
A
M
= projected spot area in the plane normal to the receiver sight line; and

HR
= fractional decrease in
w
for a path from sea level to H
R
.
A
M
is related to the target laser spot area by:
A A
M r
= cos (3.10)
Therefore, the final expression for energy density (I) at the receiver aperture for the Lambertian target is,
by substitution:
I
E
A
R
R
= (J/m
2
) (3.11)
I
BA A e
R
A
R M
R
R
R
w HR T
=
( )
2
1
(3.12)
I
G A e
R
T M
R
R
w HR T
=

( )
2
(3.13)

) (
2 2
) (
cos
) (
cos 4
R HR w
T HR w
R
R
r T
T T L
R
t
e
R
A
R D
e U
I


+
= (3.14)

| |
I
UA e
D R R
T t r
R R
L T T R
w HR T HR R
=
+
+
4
2 2 2



cos cos
( )
( )
(3.15)
If the seeker of the LGW is not turned towards the target, an additional cosine factor would be introduced
reducing the effective receiving aperture as a function of the angle between the line of sight and the
normal to the aperture (
R
). Therefore, in general:

| |
2 2 2
) (
) (
cos cos cos 4
R T T L
R R
R r t T
R R D
e UA
I
R HR T HR w



+
=
+
(3.16)
If the transmitter and receiver a collocated (case of LRF), the equation can be simplified by setting:
H
r
= H
t

HR
=
HT

r
= 0
Rr = Rt = Ro r = t
Therefore:

( )
2 2
) 2 ( 2
cos 4
o o T L
R
t T
R R D
e UA
I
HT o w



+
=

(3.17)
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 5


The term
| | ) (
R HR T HR w
R R
e
+
in Eq. (3.16) represents the two-ways atmospheric transmittance for the general
case (i.e., transmitter and receiver not collocated), denoted as
atm
in the rest of this volume. The term
) 2 (
HT o w
R
e

in Eq. (3.17) represents the two-ways transmittance for the case of transmitter-receiver
collocation (also denoted with
atm
in this volume).
The expressions derived can be used to evaluate the maximum range performance of a LRF or LTD
system, by substituting the various transmitter and receiver parameters, and solving for R
t
and R
r
. For this
purpose, the Minimum Detecatable Energy Density (MDED) at the receiver aperture is substituted for
energy density in the Eq. (3.16) or (3.17). From a practical point of view, the difficulties of this approach
for operational-level performance analysis are represented by the calculation of
atm
(a function of R
T
, R
R
,
visibility, humidity, altitude, grazing angle, etc.), the knowledge of the target characteristics (shape,
reflectivity, etc.) and, very often, the unavailability of technical data on the seeker-head detectors and
active laser systems.
Since the physical characteristics of the target are often known before performing an attack and the target is
generally extended at ranges of practical interest, it is generally sufficient to use the diffuse reflectivity of the
surface that will be illuminated, at the wavelength considered (e.g., 1.064 m). Moreover, since the
characteristics of target designators laser signals are standardised within NATO countries by the STANAG
3733, there is no much the system designer can do in order to enhance the performance of a designating
system, except than increasing the output power of the system and reducing the beam divergence. On the
other hand, some laboratory experiments (see Chapter 8 of this volume) have shown that direct measurement
of the seeker minimum detectable energy is possible, directly using the seeker and a relatively simple
instrumentation.
In most cases, it is therefore possible to estimate the performance of a LRF/LTD system as long as the
atmospheric propagation of the laser beam can be adequately modelled. This is not an easy task, especially
taking into account the considerable variation that the atmospheric parameters may experience during real
missions and for propagation paths that may exceed 10 15 km.
Additional parameters to be considered are the transmitting and receiving optics losses and the limited
integration time of the detection circuits. When the target is an extended horizontal surface, for example,
the laser can illuminate target areas whose slant-range varies significantly. This is especially true when the
laser is operating from low altitudes (i.e., low grazing angles). The result is to cause target reflections from
a given pulse transmission to be received during a relatively long time interval compared to the
transmitted pulsewidth. Receiver sensitivity, in terms of the capability of detecting a given reflected
energy, is degraded when the received pulse duration is longer than the receiver integration time. In fact,
when the detector is a peak reading threshold detector, only the energy received during an integration
period contributes effectively in achieving detection. Although the integration output does continue to rise
as long as energy is being received, the rate of rise is so slight that precise timing of the threshold crossing
becomes impossible in the presence of receiver and background noise. Accordingly, the energy received
after expiration of the integration time is useless in determining target range or performing other timing
functions. The end effect is reduced receiver sensitivity.
3.3 LASER BEAM ATMOSPHERIC PROPAGATION
Many studies have been undertaken for characterising and modelling linear and non-linear atmospheric
propagation effects on laser beams. In the following paragraphs, only a brief introduction to the
fundamentals of laser beam propagation is presented, with emphasis on those phenomena affecting the
peak irradiance at the target. Furthermore, an outline is presented of the empiric models currently used by
the Italian Air Force for PILASTER test/training operations (i.e., mission planning, safety studies and
performance analysis) with ground/airborne laser systems.
LASER SYSTEMS PERFORMANCE
3 - 6 RTO-AG-300-V26


3.3.1 Atmospheric Transmittance
Attenuation of laser radiation in the atmosphere is described by the Beers law:

( )
z
e
I
z I


= =
0
(3.18)
where is the transmittance, is the attenuation coefficient, and z is the length of the transmission path.
If the attenuation coefficient is a function of the path, then Eq. (3.18) becomes:

( )

=

z
dz z
e
0

(3.19)
The attenuation coefficient is determined by four individual processes: molecular absorption, molecular
scattering, aerosol absorption, and aerosol scattering. The atmospheric attenuation coefficient is:

a a m m
+ + + = (3.20)
where is the absorption coefficient, is the scattering coefficient, and the subscripts m and a designate the
molecular and aerosol processes, respectively. Each coefficient in Eq. (3.20) depends on the wavelength of
the laser radiation. We find it convenient at times to discuss absorption and scattering in terms of the
absorption and scattering cross sections (
a
and
s
, respectively) of the individual particles that are involved.
Thus, we can write:

a a
N = (3.21)
and also:

s s
N = (3.22)
where
a
N and
s
N are the concentrations of the absorbers and scatterers, respectively. In the absence of
precipitation, the atmosphere contains finely dispersed solid and liquid particles (of ice, dust, aromatic and
organic material) that vary in size from a cluster of a few molecules to particles of about 20 m in radius.
Particles larger than this remain airborne for a short time and are only found close to their sources. Such a
colloidal system, in which a gas (in this case, air) is the continuous medium and particles of solid or liquid
are dispersed, is known as an aerosol. Aerosol attenuation coefficients depend considerably on the
dimensions, chemical composition, and concentration of aerosol particles. These particles are generally
assumed to be homogeneous spheres that are characterized by two parameters: the radius and the index of
refraction. In general, the index of refraction is complex. Therefore, we can write:
( ) i n
n
k
i n ik n n
~
= |
.
|

\
|
= = 1 1 (3.23)
where n and k are the real and imaginary parts and = k/n is known as the extinction coefficient.
In general, both n and k are functions of the frequency of the incident radiation. The imaginary part (which
arises from a finite conductivity of the particle) is a measure of the absorption. In fact, k is referred to as
the absorption constant. It is related to the absorption coefficient of Eqs. (3.20) and (3.21) by:

c
fk

4
= (3.24)

where c is the speed of light in a vacuum and f is the frequency of the incident radiation. For the wavelength
range of greater interest in laser beam propagation (the visible region to about 15 m) the principal
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 7


atmospheric absorbers are the molecules of water, carbon dioxide, and ozone. Attenuation occurs because
these molecules selectively absorb radiation by changing vibrational and rotational energy states. The two
gases present in greatest abundance in the earths atmosphere, nitrogen (
2
N ) and oxygen (
2
O ),
are homonuclear, which means that they possess no electric dipole moment and therefore do not exhibit
molecular absorption bands. The atmospheric spectral transmittance (%) measured over a 1820-m
horizontal path at sea level is shown in Figure 3-2. The molecule responsible for each absorption band is
shown in the upper part of the figure. It is evident that O H
2
and
2
CO are by far the most important
absorbing molecules. This is also the case for the range of altitudes extending from sea level to about 12 km.
Depending on weather conditions, altitude, and geographical location, the concentration of O H
2
varies
between 10
-3
and 1 percent (by volume). The concentration of
2
CO varies between 0.03 and 0.04 percent.
Other absorbing molecules found in the atmosphere are methane (
4
CH ), with a concentration of around
1.5 10
-4
percent; nitrous oxide ( O N
2
), with a concentration of around 3.5 10
-5
percent; carbon monoxide
(CO) with a typical concentration of 2 10
-5
percent; and ozone (
3
O ), with a concentration as large as 10
-3

percent at an altitude of around 30 km. The concentration of ozone near sea level is negligible. In Figure 3-2
the wavelength intervals where the transmittance is relatively high are called atmospheric windows.


























Figure 3-2: Sea-Level Transmittance Over a 1820 m Horizontal Path [3].
Obviously, for efficient energy transmission the laser wavelength should fall well within one of these
windows. There are a total of eight such windows within the wavelength range extending from 0.72 to
15.0 m. The window boundaries are listed in Table 3-1.


Far Infrared

Near Infrared
VIII VII VI V
IV
III
II

Mid Infrared

I
LASER SYSTEMS PERFORMANCE
3 - 8 RTO-AG-300-V26


Table 3-1: Wavelength Regions of Atmospheric Windows
Window Number Window Boundaries (m)
I 0.72 0.94
II 0.94 1.13
III 1.13 1.38
IV 1.38 1.90
V 1.90 2.70
VI 2.70 4.30
VII 4.30 6.00
VIII 6.00 15.0
The scattering coefficient in Eqs. (3.20) and (3.22) also depends on the frequency of the incident radiation
as well as the index of refraction and radius of the scattering particle. The incident electromagnetic wave,
which is assumed to be a plane wave in a given polarization state, produces forced oscillations of the bound
and free charges within the sphere. These oscillating charges in turn produce secondary fields internal and
external to the sphere. The resulting field at any point is the vector sum of the primary (plane wave) and
secondary fields. Once the resultant field has been determined, the scattering cross section is obtained from
the following relationship:

vector poynting incident averaged time the of magnitude
scatterer by scattered power total
s

= (3.25)
In the scattering process there is no loss of energy but only a directional redistribution which may lead to a
significant reduction in beam intensity for large path lengths. As is indicated in Table 3-2, the physical
size of the scatterer determines the type of scattering. Thus, air molecules that are typically several
angstrom units in diameter lead to Rayleigh scattering, whereas the aerosols scatter light in accordance
with the Mie theory. Furthermore, when the scatterers are relatively large, such as the water droplets found
in fog, clouds, rain, or snow, the scattering process is more properly described by diffraction theory.
Table 3-2: Types of Atmospheric Scattering
Type of Scattering Size of Scatterer
Rayleigh Scattering Larger than electron but smaller than
Mie Scattering Comparable in size to
Non-selective Scattering Much larger than
3.3.2 Computer Codes
In principle, one could determine the exact composition of the atmosphere over the path of interest and,
employing the physics of molecular and aerosol extinction, compute the atmospheric extinction
coefficient. Because of the wide variations in weather conditions and sparsity of data on some atmospheric
constituents, it is desirable to adopt an engineering approach to atmospheric modelling. The required
model should include several weather conditions and should be validated with laboratory and field data.
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 9


To deal with these complex phenomena, the Phillips Laboratory of the Geophysics Directorate at
Hanscom Air Force Base (Massachusetts) has developed codes to predict transmittance/radiance effects for
varying conditions. Particularly, they have created LOWTRAN (LOW spectral resolution TRANsmission
code), FASCODE (FASt atmospheric signature CODE), MODTRAN (MODerate spectral resolution
TRANsmission code), and HITRAN (HIgh resolution TRANsmission code). Furthermore, in recent years,
powerful tools for the assessment and exploitation of propagation conditions together with range
performance models for military systems have become available.
It is impossible to present in a fully comprehensive way all available tools. Instead, some relevant
information is given in Ref. [1]-[3]. In the following paragraphs, only the empirical models selected for
the initial versions of the PILASTER Mission Planning and Analysis (MPA) software tools are described.
3.3.3 Elder-Strong-Langer (ESL) Model for
ai

A simple approach, yielding approximate values of the absorption coefficient, has been suggested by Elder
and Strong [4] and modified by Langer [5]. Their approach is particularly useful because it provides a
means of relating the atmospheric transmission of the i
th
window to the relative humidity (i.e., a readily
measurable parameter). The assumption is that variations in the transmission are caused by changes in the
water content of the air. Specifically, changes in the concentration of H
2
O cause changes in the absorption,
and changes in the size and number of water droplets with humidity cause changes in the scattered
component. This is a valid assumption since the other atmospheric constituents have a reasonably constant
effect on the transmittance of a given atmospheric window.
It is customary to express the number of H
2
O molecules encountered by the beam of light in terms of the
number of precipitable millimetres of water in the path. Specifically, the depth of the layer of water that
would be formed if all the water molecules along the propagation path were condensed in a container
having the same cross-sectional area as the beam is the amount of precipitable water. A cubic meter of air
having an absolute humidity of grams per m
3
would yield condensed water that cover a l m
2
area and
have a depth of:

3
10

= w (3.26)
w is the precipitable water having units of mm per meter of path length. For a path length of z meters
Eq. (3.26) becomes:
z w =

3
10 (3.27)
where w is now the total precipitable water in millimetres. The value of , the density of water vapour, can
be found by multiplying the appropriate number in Table 3-3 by the relative humidity (RH).
LASER SYSTEMS PERFORMANCE
3 - 10 RTO-AG-300-V26


Table 3-3: Mass of Water Vapour in Saturated Air (g/m
3
)
Temperature
(C) 0 1 2 3 4 5 6 7 8 9
-20 0.89 0.81 0.74 0.67 0.61 0.65
-10 2.15 1.98 1.81 1.66 1.52 1.40 1.28 1.18 1.08 0.98
-0 4.84 4.47 4.13 3.81 3.52 3.24 2.99 2.75 2.54 2.34
0 4.84 5.18 5.54 5.92 6.33 6.76 7.22 7.70 8.22 8.76
10 9.33 9.94 10.57 11.25 11.96 12.71 13.50 14.34 15.22 16.17
20 17.22 18.14 19.22 20.36 21.55 22.80 24.11 25.49 27.00 28.45
30 30.04 31.70 33.45 35.28 37.19 39.19
Similar numerical results can be obtained using the following equation [6], which is convenient for
computer code implementations:

( )
(

|
.
|

\
|


=
16 . 273
ln 31 . 5
16 . 273 22 . 25
exp 8 . 1322
T
T
T
T
RH
(3.28)
where RH is the relative humidity (as a fraction), and T is the absolute temperature (K).
Based on the work done by Elder and Strong [4], two empirical expressions, developed by Langer [5],
can be used to calculate the absorptive transmittance
ai
for the i
th
window for any given value of the
precipitable water content. These expressions are:

w A
ai
i
e

= , for
i
w w < (3.29)

i
w
w
k
i
i ai

|
.
|

\
|
= , for
i
w w > (3.30)
where A
i
, k
i
,
i
and w
i
are constants whose values for each atmospheric window are listed in Table 3-4.
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 11


Table 3-4: Constants to be Used in Eqs. (3.34) and (3.35)
Constants

Window

A
i

k
i

i

w
i
I 0.0305 0.800 0.112 54
II 0.0363 0.765 0.134 54
III 0.1303 0.830 0.093 2.0
IV 0.211 0.802 0.111 1.1
V 0.350 0.814 0.1035 0.35
VI 0.373 0.827 0.095 0.26
VII 0.598 0.784 0.122 0.165
In summary, Eqs. (3.29) and (3.30), together with Eq. (3.27) and Table 3-3 (or Eq. 3.28), provide
information that can be used to obtain an estimate of the absorptive transmittance (
ai
) of laser beams
having wavelengths that fall within the various atmospheric windows. The results apply to horizontal
paths in the atmosphere near sea level and for varying relative humidity. To obtain the total atmospheric
transmittance we must multiply
ai
by
si
(i.e., the transmittance due to scattering only).
3.3.4 Empirical Expressions for
si

Based on rigorous mathematical approaches, the scattering properties of the atmosphere due to the aerosol
particles are difficult to quantify, and it is difficult to obtain an analytic expression for the scattering
coefficient that will yield accurate values over a wide variety of conditions. However, an empirical
relationship that is often used to model the scattering coefficient [7] has the form:
( )
4
2 1

+ =

C C (3.31)
where C
1
, C
2
, and are constants determined by the aerosol concentration and size distribution, and is
the wavelength of the radiation. The second term accounts for Rayleigh scattering. Since for all
wavelengths longer than about 0.3 m the second term is considerably less than the first, it may be
neglected. It has been found that

3 0 3 1 . . produces reasonable results when applied to aerosols with
a range of particle sizes.
An attempt has also been made to relate and C
1
to the meteorological range. The apparent contrast C
z
, of
a source when viewed at = 0.55 m from a distance z is by definition:

bz
bz sz
z
R
R R
C

= (3.32)
Where R
sz
and R
bz
are the apparent radiances of the source and its background as seen from a distance z.
For m 55 . 0 = , the distance at which the ratio:
LASER SYSTEMS PERFORMANCE
3 - 12 RTO-AG-300-V26


02 . 0
0
0 0
0
=

= =
b
b s
bz
bz sz
z
R
R R
R
R R
C
C
V (3.33)
is defined as the meteorological range V (or visual range). It must be observed that this quantity is
different from the standard observer visibility (V
obs
). Observer visibility is the greatest distance at which it
is just possible to see and identify a target with the unaided eye. In daytime, the object used for V
obs

measurements is dark against the horizon sky (e.g., high contrast target), while during night time the target
is a moderately intense light source. The International Visibility Code (IVC) is given in Table 3-5. It is
evident that, while the range of values for each category is appropriate for general purposes, it is too broad
for scientific applications.
Table 3-5: International Visibility Code (IVC)
DESIGNATION VISIBILITY
Dense Fog 0 50 m
Thick Fog 50 200 m
Moderate Fog 200 500 m
Light Fog 500 1 km
Thin Fog 1 2 km
Haze 2 4 km
Light Haze 4 10 km
Clear 10 20 km
Very Clear 20 50 km
Exceptionally Clear > 50 km
Visibility is a subjective measurement estimated by a trained observer and as such can have large
variability associated with the reported value. Variations are created by observers having different
threshold contrasts looking at non-ideal targets. Obviously, visibility depends on the aerosol distribution
and it is very sensitive to the local meteorological conditions. It is also dependent upon the view angle
with respect to the sun. As the sun angle approaches the view angle, forward scattering into the line-of-
sight increases and the visibility decreases. Therefore, reports from local weather stations may or may not
represent the actual conditions at which the experiment is taking place. Since meteorogical range is
defined quantitatively using the apparent contrast of a source (or the apparent radiances of the source and
its background) as seen from a certain distance, it eliminates the subjective nature of the observer and the
distinction between day and night. Unfortunately, carelessness has often resulted in using the term
visibility when meteorological range is meant. To insure that there is no confusion, observer-visibility
(V
obs
) will be used in this volume to indicate that it is an estimate.
If only V
obs
is available, the meteorological range (V) can be estimated [6] from:
( )
obs
V . . V 3 0 3 1 (3.34)
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 13


From Eq. (3.33), if we assume that the source radiance is much greater than the background radiance
(i.e., R
s
>> R
b
) and that the background radiance is constant (i.e., R
bo
= R
bz
), then the transmittance at
= 0.55 m (where absorption is negligible) is given by:
02 . 0
0
= =
V
s
sv
e
R
R

(3.35)
Hence, we have:
91 . 3 ln
0
= =
|
|
.
|

\
|
V
R
R
s
sv
(3.36)
and also:



= =
1
91 . 3
C
V
(3.37)
It follows from Eq. (3.36) that the constant C
1
is given by:

55 0
91 3
1
.
V
.
C = (3.38)
With this result the transmittance at the centre of the i
th
window is:

z
. V
.
si
i
e

|
|
.
|

\
|


=

55 0
91 3
(3.39)
where
i
must be expressed in microns.
If, because of haze, the meteorological range is less than 6 km, the exponent is related to the
meteorological range by the following empirical formula:

3
585 . 0 V = (3.40)
where V is in kilometres. When V 6 km, the exponent can be calculated by:
025 1 0057 0 . V . + = (3.41)
For exceptionally good visibility = 1.6, and for average visibility 1.3. In summary, Eq. (3.39),
together with the appropriate value for , permits us to compute the scattering transmittance at the centre
of the i
th
window for any propagation path, if the meteorological range V is known. It is important to note
here that in general the transmittance will, of course, also be affected by atmospheric absorption, which
depending on the relative humidity and temperature may be larger than
si
3.3.5 Propagation Through Haze and Precipitation
Haze refers to the small particles suspended in the air. These particles consist of microscopic salt crystals,
very fine dust, and combustion products. Their radii are less than 0.5 m. During periods of high
humidity, water molecules condense onto these particles, which then increase in size. It is essential that
these condensation nuclei be available before condensation can take place. Since salt is quite hygroscopic,
it is by far the most important condensation nucleus. Fog occurs when the condensation nuclei grow into
water droplets or ice crystals with radii exceeding 0.5 m. Clouds are formed in the same way; the only
distinction between fog and clouds is that one touches the ground while the other does not. By convention
fog limits the visibility to less than 1 km, whereas in a mist the visibility is greater than 1 km.
LASER SYSTEMS PERFORMANCE
3 - 14 RTO-AG-300-V26


We know that in the early stages of droplet growth the Mie attenuation factor K depends strongly on the
wavelength. When the drop has reached a radius a 10 the value of K approaches 2, and the scattering is
now independent of wavelength, i.e., it is non-selective. Since most of the fog droplets have radii ranging
from 5 to 15 m they are comparable in size to the wavelength of infrared radiation. Consequently the
value of the scattering cross section is near its maximum. It follows that the transmission of fogs in either
the visible or IR spectral region is poor for any reasonable path length. This of course also applies to
clouds.
Since haze particles are usually less than 0.5 m, we note that for laser beams in the IR spectral region
1 << a and the scattering is not an important attenuation mechanism. This explains why photographs of
distant objects are sometimes made with infrared-sensitive film that responds to wavelengths out to about
0.85 m. At this wavelength the transmittance of a light haze is about twice that at 0.5 m. Raindrops are
of course many times larger than the wavelengths of laser beams. As a result there is no wavelength-
dependent scattering. The scattering coefficient does, however, depend strongly on the size of the drop.
Middleton [7],[8] has shown that the scattering coefficient with rain is given by:

3
6
10 25 1
a
t x
.
rain


= (3.42)
where x/t is the rainfall rate in centimetres of depth per second and a is the radius of the drops in
centimetres. Rainfall rates for four different rain conditions and the corresponding transmittance (due to
scattering only ) of a 1.8 km path are shown in Table 3-6 [9]. These data are useful for order of magnitude
estimates. In order to obtain accurate estimates, the concentrations of the different types of rain drops
(radius) and the associated rainfall rates should be known. In this case, the scattering coefficient can be
calculated as the sum of the partial coefficients associated to the various rain drops.
Table 3-6: Transmittance of a 1.8 km Path Through Rain
Rainfall (cm/h) Transmittance (1.8 km path)
0.25 0.88
1.25 0.74
2.5 0.65
10.0 0.38
A simpler approach, used in LOWTRAN, gives good approximations of the results obtained with
Eq. (3.42) for most concentrations of different rain particles. Particularly, in LOWTRAN, the scattering
coefficient with rain has been empirically related only to the rainfall rate t x (expressed in mm/hour),
as follows [6]:

63 0
365 0
.
rain
t
x
. |
.
|

\
|

(3.43)
Table 3-7 provides representative rainfall rates which can be used in Eqs. (3.42) and (3.43), when no direct
measurements are available, to obtain order of magnitude estimations of
rain
[10].
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 15


Table 3-7: Representative Rainfall Rates
Rain Intensity Rainfall (mm/hour)
Mist 0.025
Drizzle 0.25
Light 1.0
Moderate 4.0
Heavy 16
Thundershower 40
Cloud-Burst 100
In the presence of rain, in addition to the scattering losses calculated with Eq. (3.42) or (3.43), there are,
of course, losses by absorption along the path, and these must be included in the calculation of the total
atmospheric transmittance with rain.
3.3.6 PILASTER Combined Model
Combining the equations presented in Sections 3.2.2, 3.2.3 and 3.2.4, the set of equations presented in
Table 3-8 were obtained, for calculating the atmospheric transmittance (
atm
) in the various conditions,
with transmitter and receiver collocated.
Table 3-8: Transmittance Equations for Transmitter and Receiver Collocated
Case Cond. Equations N
A
V 6 km
w>wi

(0.0057
.
V+1.025)
55 0
91 3
i i
. V
.
z
i
i atm
e
w
w
k
.
|

\
|


.
|

\
|
=




(3.44)

B
V 6 km
w<wi

( )
(
(

|
.
|

\
|
+
+
=
025 1 0057 0
55 0
91 3
. V .
i
i
. V
.
w A z
atm
e



(3.45)

C
V < 6 km
w<wi

(
(

|
.
|

\
|
+

=
3
585 0
55 0
91 3
V .
i
i
. V
.
w A z
atm
e



(3.46)

D
V < 6 km
w>wi

3
585 0
55 0
91 3
V .
i i
. V
.
z
i
i atm
e
w
w
k

|
.
|

\
|

|
.
|

\
|
=



(3.47)

R
1

Rain
w<w
i


(
(

|
.
|

\
|

=
63 0
365 0
.
i
t
x
. z
w A
atm
e e



(3.48)

R
2

Rain
w>wi
(
(

|
.
|

\
|

|
.
|

\
|
=
63 0
365 0
.
i
t
x
. z
i
i atm
e
w
w
k



(3.49)

LASER SYSTEMS PERFORMANCE
3 - 16 RTO-AG-300-V26


The cases R
1
and R
2
in Table 3-8 are independent of meteorological range (V). Straightforward numerical
analysis shows that the
atm
estimates obtained with rain using Eqs. (3.48) and (3.49), are always less than
the corresponding transmittance estimates obtained with Eqs. (3.46) and (3.47) with dry-air conditions and
V < 6 km, for rainfall rates 1 t x (i.e., from light rain to cloud-burst).
In the case of transmitter and receiver not collocated (e.g., LTD/LGW combination), the equations in
Table 3-8 have to be modified, taking into account that the total laser path (z) is given by the sum of the
range transmitter-target and target-receiver (see Figure 3-1). Therefore, we have:

r t
R R z + = (3.50)
Denoting with the subscripts t and r the terms relative to the transmitting and receiving paths respectively,
we have that the total atmospheric transmittance (
tot
) is given by:

r t tot
= (3.51)
Therefore, in order to account for all possible cases, we have to consider the 2
3
possible combinations
referring to dry-air ( km 6 km 6 < V V ,
i t i t
w w w w < and
i r i r
w w w w < ), and the 2
2

combinations relative to rainy conditions (
i t i t
w w w w < and
i r i r
w w w w < ).
It should be considered, however, that the condition
i t
w w < is not likely to occur in many cases of
practical interest with LTD/LGW systems. From Eq. (3.27), we obtain the maximum transmitter distance
(R
max
) from which the condition
i t
w w < is verified:

3
10 <

i
max
w
R (3.52)
In normal dry-air conditions (e.g., T = 24C and RH = 75%) R
max
equates to about 3 km. This is a distance
very short in many real operational scenarios. Obviously, whit rainy conditions, the range R
max
would be
even shorter. Table 3-9 and Table 3-10 show the equations developed for all dry-air and rain cases
considered.
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 17


Table 3-9: ESLM-Dry Equations for Transmitter and Receiver Not Collocated
Case Cond.
Equations
n
E
V 6 km
w
t
w
i
w
r
w
i




( )
t
. V .
i i
R
. V
.
t
i
i
e
w
w
k
025 1 0057 0
55 0
91 3
+
|
|
.
|

\
|

|
|
.
|

\
|

( )
r
. V .
i i
R
. V
.
r
i
i
e
w
w
k
025 1 0057 0
55 0
91 3
+
|
|
.
|

\
|

|
|
.
|

\
|




(3.53)

F
V 6 km
w
t
w
i
w
r
< w
i

( )
t
. V .
i i
R
. V
.
t
i
i
e
w
w
k
025 1 0057 0
55 0
91 3
+
|
|
.
|

\
|

|
|
.
|

\
|

( )
r
. V .
i
r i
R
. V
.
w A
e
025 1 0057 0
55 0
91 3
+
|
|
.
|

\
|





(3.54)

G
V < 6 km
w
t
w
i
w
r
< w
i

t
V
i
i
R
V
t
i
i
e
w
w
k
3
585 . 0
55 . 0
91 . 3

|
.
|

\
|

|
|
.
|

\
|

r
V
i
r i
R
V
w A
e
3
585 . 0
55 . 0
91 . 3

|
.
|

\
|





(3.55)

H
V < 6 km
w
t
w
i
w
r
w
i

t
V
i i
R
V
t
i
i
e
w
w
k
3
585 . 0
55 . 0
91 . 3

|
.
|

\
|

|
|
.
|

\
|

r
V
i i
R
V
r
i
i
e
w
w
k
3
585 . 0
55 . 0
91 . 3

|
.
|

\
|

|
|
.
|

\
|




(3.56)

I
V 6 km
w
t
< w
i
w
r
w
i

( )
t
. V .
i
t i
R
. V
.
w A
e
025 1 0057 0
55 0
91 3
+
|
.
|

\
|


( )
r
. V .
i
i
R
. V
.
r
i
i
e
w
w
k
025 1 0057 0
55 0
91 3
+
|
.
|

\
|

|
|
.
|

\
|



(3.57)

J
V 6 km
w
t
< w
i
w
r
< w
i

( )
t
. V .
i
t i
R
. V
.
w A
e
025 1 0057 0
55 0
91 3
+
|
.
|

\
|



( )
r
. V .
i
r i
R
. V
.
w A
e
025 1 0057 0
55 0
91 3
+
|
|
.
|

\
|





(3.58)

K
V < 6 km
w
t
< w
i
w
r
< w
i

t
V .
i
t i
R
. V
.
w A
e
3
585 0
55 0
91 3

|
.
|

\
|



r
V .
i
r i
R
. V
.
w A
e
3
585 0
55 0
91 3

|
|
.
|

\
|





(3.59)

L
V < 6 km
w
t
< w
i
w
r
w
i

t
V .
i
t i
R
. V
.
w A
e
3
585 0
55 0
91 3

|
.
|

\
|



r
V .
i i
R
. V
.
r
i
i
e
w
w
k
3
585 0
55 0
91 3

|
|
.
|

\
|

|
|
.
|

\
|





(3.60)


( )
( )
r t
. V .
i
i
R R
. V
.
r t
i
i
e
w w
w
k
+ |
.
|

\
|

+
|
|
.
|

\
|
025 1 0057 0
55 0
91 3
2
2

( )
( )
r t
. V .
i
r i
i
R R
. V
.
w A
t
i
i
e
w
w
k
+ |
.
|

\
|

+
|
|
.
|

\
|
025 1 0057 0
55 0
91 3

( )
r t
V .
i
r i
i
R R
. V
.
w A
t
i
i
e
w
w
k
+ |
.
|

\
|


|
|
.
|

\
|
3
585 0
55 0
91 3

( )
r t
V .
i
i
R R
. V
.
r t
i
i
e
w w
w
k
+ |
.
|

\
|


|
|
.
|

\
|
3
585 0
55 0
91 3
2
2

( )
( )
r t
. V .
i
t i
i
R R
. V
.
w A
r
i
i
e
w
w
k
+ |
.
|

\
|

+
|
|
.
|

\
|
025 1 0057 0
55 0
91 3

( )
( )
r t
. V .
i
r i t i
R R
. V
.
w A w A
e
+ |
.
|

\
|

+ 025 1 0057 0
55 0
91 3
( )
r t
V .
i
r i t i
R R
. V
.
w A w A
e
+ |
.
|

\
|


3
585 0
55 0
91 3
( )
r t
V .
i
t i
i
R R
. V
.
w A
r
i
i
e
w
w
k
+ |
.
|

\
|


|
|
.
|

\
|
3
585 0
55 0
91 3

LASER SYSTEMS PERFORMANCE


3 - 18 RTO-AG-300-V26


Table 3-10: ESLM-Rain Equations for Transmitter and Receiver Not Collocated
Case Cond.
Equations
N
R
3

Rain
w
t
w
i
w
r
w
i




t
.
i
R
t
x
.
t
i
i
e
w
w
k
63 0
365 0 |
.
|

\
|

|
|
.
|

\
|

r
.
i
R
t
x
,
r
i
i
e
w
w
k
63 0
365 0 |
.
|

\
|

|
|
.
|

\
|




(3.61)


R
4

Rain
w
t
w
i
w
r
< w
i

t
.
i
R
t
x
.
t
i
i
e
w
w
k
63 0
365 0 |
.
|

\
|

|
|
.
|

\
|

r
.
r i
R
t
x
. w A
e
63 0
365 0 |
.
|

\
|




(3.62)

R
5

Rain
w
t
< w
i
w
r
w
i

t
.
t i
R
t
x
. w A
e
63 0
365 0 |
.
|

\
|

r
.
i
R
t
x
.
r
i
i
e
w
w
k
63 0
365 0 |
.
|

\
|

|
|
.
|

\
|




(3.63)

R
6

Rain
w
t
< w
i
w
r
< w
i



t
.
t i
R
t
x
. w A
e
63 0
365 0 |
.
|

\
|


r
.
r i
R
t
x
. w A
e
63 0
365 0 |
.
|

\
|





(3.64)
The equations presented in the Table 3-8, Table 3-9 and Table 3-10 represent the combined Elder-Strong-
Langer-Middleton (ESLM) model, relative to laser beam horizontal-path propagation at sea-level both in
dry-air and rain conditions. The validation process of the ESLM model, before incorporation in the
PILASTER MPA tools, was undertaken during this research using experimental data collected during
ground trials. Furthermore, corrections to be applied with increasing altitudes and with various laser slant-
path grazing angles were determined using data collected in flight tests. The results of these activities are
described in the Chapters 8 and 9 of this volume.
3.3.7 Refractive Index Variations
When a laser beam passes through air, the randomly fluctuating air temperature produces small density
and refractive index inhomogeneities that affect the beam in at least three different ways. Considering for
example an initially well-defined phase front propagating through a region of atmospheric turbulence.
Because of random fluctuations in phase velocity the initially well defined phase front will become
distorted. This alters and redirects the flow of energy in the beam. As the distorted phase front progresses,
random changes in beam direction (Beam Wander) and intensity fluctuations (Scintillation) occur.
The beam is also found to spread in size beyond the dimensions predicted by diffraction theory.
The cause of all this, as we have stated, is atmospheric turbulence that arises when air parcels of different
temperatures are mixed by wind and convection. The individual air parcels, or turbulence cells, break up
into smaller cells and eventually lose their identity. In the meantime, however, the mixing produces
fluctuations in the density and therefore in the refractive index of air. To describe these random processes,
one must have a way of defining the fluctuations that are characteristic of turbulence. The most common
approaches adopted may be found in Strohbehn [12] and Weichel [3].
( )
r t
.
i
R R
t
x
.
r t
i
i
e
w w
w
k
+ |
.
|

\
|

|
|
.
|

\
|
63 0
365 0
2
2

( )
r t
.
r i
i
R R
t
x
. w A
t
i
i
e
w
w
k
+ |
.
|

\
|

|
|
.
|

\
|
63 0
365 0

( )
r t
.
t i
i
R R
t
x
. w A
r
i
i
e
w
w
k
+ |
.
|

\
|

|
|
.
|

\
|
63 0
365 0

( ) ( )
r t
.
r t i
R R
t
x
. w w A
e
+ |
.
|

\
|

63 0
365 0

LASER SYSTEMS PERFORMANCE


RTO-AG-300-V26 3 - 19


3.3.8 Other Propagation Effects
The propagation of a laser beam through atmospheric turbulence is a linear phenomenon in that the air is
not affected by the beam. Strictly speaking, this is only true for beams of relatively low irradiance. As the
beam irradiance is increased, molecular absorption will lead to temperature gradients in the medium that
in turn induce density and index-of-refraction changes. The final result is a medium whose optical
properties have been altered. This phenomenon is non-linear, in that the beam irradiance distribution leads
to index-of-refraction changes, which in turn alter the beams irradiance distribution, which alters the
refractive index, etc.
Non-linear propagation effects typically include: Thermal Blooming (whose consequence is that the
divergence angle is considerably more than that due to diffraction alone), Kinetic Cooling (resulting in a
temporary focusing effect and less than diffraction limited beam spreading), and Bleaching (1 5 sec
duration pulses may under certain conditions saturate the absorption mechanism and thereby reduce the
atmospheric transmittance). Also aerodynamic effects influence the performance of the airborne systems.
These effects can be grouped in two categories:
Aeromechanical Effects, arising from interactions of the external flow field with the airborne
platform. This base motion, in concert with intrinsic platform sources of vibration (e.g., engines,
pumps, fluid flow), defines the overall mechanical jitter environment in which the laser system
must operate. Jitter can result in spurious laser beam motion on target, as well as general
misalignment of optical elements.
Aero-Optical (AO) Effects: These are caused by refraction index variations induced by the
platform moving through the flow field. This results in reduced far-field peak intensity as well as
beam spread and wander for outgoing wave fronts (for imaging systems, these several effects
manifest themselves as loss of contrast and resolution).
An outline of these additional propagation effects can be found in Ref. [13].
3.4 LASER SCATTERING AND TARGET CROSS SECTION
The scattering and propagation of laser light obey the same set of laws as radio frequency waves, that is,
those set forth by Maxwells equations and the boundary conditions. However, the wavelength of laser
light is so small that minute particles and even molecules represent significant scatterers. Target surfaces
are generally very rough at laser wavelengths and, consequently, the random or diffuse reflection
component frequently dominates. In fact, there may not be any significant specular component to the laser
cross section, in many cases. Sometimes, however, significant specular reflections and retro-reflections
(opposition effects) are observed from certain target surfaces. Furthermore, in general, the overall
scattering pattern produced by a certain (complex) target illuminated by a laser beam shows a marked
dependency on the illumination incidence angle.
When examining the diffuse reflection component, the maximum amount of reflected energy is reflected
90 (normal) to the surface - independent of the incoming beam angle of arrival, and the energy falls off as
a function of the cosine of the angle off of surface normal.
A surface that is a perfect diffuser scatters incident light equally in all directions. For such an ideal
surface, the intensity (W/m
2
) of diffusely reflected light is given by:
cos
d i d
k I I = with
(

2
, 0

(3.65)
LASER SYSTEMS PERFORMANCE
3 - 20 RTO-AG-300-V26


where I
i
is the intensity of the light source at the target, is the angle between the surface normal and a
line from the surface illuminated point to the light source (considered as a point source). The constant k
d
is
the diffuse reflectivity, which depends on the nature of the material and the wavelength of the incident
light. Eq. (3.70) may be also expressed in the vector form:
( ) N L k I I
d i d

= (3.66)
where L

and N

are the vectors illustrated in Figure 3-3.
























Figure 3-3: Reflection Geometry.
As described before, any reflection from a practical surface should be considered as (at least) the sum of a
specular component and a diffuse component. The existence of these two component has been shown
experimentally and is not a consequence of choice of a particular model. A surface attribute that is
important to model is the surface roughness. A perfectly smooth surface reflects incident radiation in a
single direction. A rough surface tends to scatter incident radiation in every direction, although certain
directions may contain more reflected energy than others. This behaviour is obviously also dependent on
the wavelength of radiation; a surface that is smooth for certain wavelengths may be rough for others.
For example, oxidised or unpolished metal is smooth for radio waves ( = 10
-2
m) and rough for radiation
in the near-infrared (NIR) part of the spectrum. In general, metals can be prevalently diffuse or specular
reflectors in the NIR depending on whether they are polished or not. So reflection is not only dependent on
the material but also on its surface properties. Another factor in reflection in the grazing angle of the
incident laser source. This can in fact determine the entity of the overall reflected signal and of the two
reflection components.
Therefore, a realistic reflection model should at least represent the target surface as some combination of a
perfect diffuse reflector and a perfect specular surface. One of the earlier and still quite popular models is the
Phong model [14]. This model can be used for fitting the results of experimental bi-directional reflectivity
measurements and for computer simulation programs. In the Phong model, the bi-directional spectral
reflectivity is expressed by:
(Source direction)
(Specular direction)
(Viewer direction)
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 21


n
spec diff
'
cos k k + = (3.67)
where k
diff
is the fraction of energy diffusely reflected and k
spec
is the fraction specularly reflected. The model
can be given in terms of the unit vectors associated with the geometry of the point under consideration
(Figure 3-3). Therefore, for the reflected intensity, we may write:
( ) ( ) | | A cos k cos k I I
n
s d i
+ + = (3.68)
( ) ( ) | | A V R k N L k I I
n
s d i
+ + = (3.69)
where k
s
is the specular reflection coefficient (a function of the material characteristics and incidence
angle), n is the index that controls the dimensions of the specular highlight, and A is an additional term
accounting for reflection of sunlight at the wavelength considered (day-time operations). This can be also
modelled as:
( ) ( ) | |
' n
s
'
d
cos k cos k E A

+ = (3.70)
where E

is the solar spectral irradiance at the wavelength of the laser , and is the angle between the
solar illumination and the normal to the target reflecting surface.
Figure 3-4 shows the variation in light intensity at a point P on a surface calculated using Eq. (3.69).
The intensity variation is shown as a profile (i.e., a function of the orientation of V). The intensity at P is
given by the length of V from P to its intersection with the profile. The semicircular part of the profile is
the contribution from the diffuse term. The specular part of the profile is shown for different values of the
index n.

















Figure 3-4: Intensity as a Function of V Orientation (with Different Values of n).
Note that, in general, the higher is the value of n, the tighter is the specular highlight. Figure 3-5 shows
the resulting combinations of the two reflection components, obtained by keeping fixed the value of n
(e.g., n = 100) and varying the angle .

P
L
N
n increase
R
LASER SYSTEMS PERFORMANCE
3 - 22 RTO-AG-300-V26








Figure 3-5: Reflection Components with Various Angles.
Figure 3-6 shows a typical surface which contains both specular and diffuse reflections with a 55%
specular component and a 45% diffuse component ( = 50, n = 100).





















Figure 3-6: Specular and Diffuse Reflection Components.
In most practical cases with LTD/LGW systems, the diffuse component alone is assumed when describing
target reflectivity, since the diffuse reflection component is what the weapon will have the highest
probability of tracking during flight. Typical diffuse reflectivity values at = 1.064 m are listed in
N
Incident
Beam
Specular
Reflection
Diffuse
Reflection
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 23


Table 3-11. It is worth to notice that glass, water and highly polished surfaces are poor surfaces to
designate since they reflect most of the laser energy back along one direction only (i.e., they are specular
reflectors).
Table 3-11: Approximate Reflectivity at = 1.064 m
Material Diffuse Reflectivity
Matt Black Paint 4 15%
Dirty Olive Drab Paint 5 15%
Soil 15 25%
Brick 15 65%
Vegetation (Glossy Foliage) 30 70%
Asphalt 10 25%
Concrete 10 40%
IR Reflecting Paint 30 55%
3.5 LTD/LGW OPERATIONAL CONSIDERATIONS
Global requirements for mission planning with a particular laser designation system may be initially
established by examining the LTD and LGW operating slant-ranges required to successfully perform the
mission (e.g., optimal delivery of a particular laser weapon). These ranges may vary from a few hundred
feet for a ground designator to over 100,000 feet for operational delivery of a Paveway III LGB. Thus,
mission planning with a particular LTD system must have an operational input that factors in the slant-
ranges expected for various types of delivery tactics. Mission planning to determine the optimal weapon
release point involves a number of factors, including the post-release designation manoeuvre to be
employed, the maximum slant-range at weapon impact, the target size, laser system error budget, laser
power, etc. What follows is a discussion of the primary factors necessary for determining the optimal
release range.
3.5.1 Target Size
Target dimensions are a critical factor in LTD/LGW mission planning. These dimensions, along with the
slant-range requirements must then be factored together with the characteristics of the designator.
In addition, it must be remembered that designation tactics will generally reduce the apparent target size
by varying degrees due to the oblique perspective most manoeuvres will generate.
As an example, if a weapon can achieve a 10 feet Circular Error Probability (CEP), then it is appropriate
that the designator aiming capability must equal or exceed that requirement in order to meet a suitable
weapon impact criteria for the weapon. As an example problem, a hardened shelter access cover, roughly
20 feet in diameter, will be used as a target. This target dimension equates to a 10 feet CEP where
50 percent of our hypothetical weapon releases should fall on the target face. Thus, one must see and
identify this target from the desired vantage point and also be able to maintain the laser energy on the
target from release to impact. Weapon system error sources challenge this ability to keep the spot on the
target as described below.
LASER SYSTEMS PERFORMANCE
3 - 24 RTO-AG-300-V26


3.5.2 LTD Systems Error Sources and Effects
Error sources such as laser spot spillover, boresight errors, jitter, and tracking errors, cause large reductions
in LGW delivery effectiveness. The following is a discussion of the most common error sources in laser
designator systems and the effects of these errors on designation performance.
3.5.2.1 Laser Spot Spillover
Several characteristics of the laser beam must be tightly controlled if the beam is to be maintained on the
desired target surface. First, the laser beam spot should be smaller than the target face. As the LTD
produces a beam that diverges as it propagates along the path between the laser and the target, beam
spillover effects often degrade weapon accuracy both when designation is performed by a ground LTD or
an airborne LTD (see Figure 3-7).







Figure 3-7: Laser Spot Spillover.
Laser beam divergence should therefore be accounted, and appropriate terminal slant-ranges and grazing
angles should be chosen such that the spot elongation will not cause spillover around the target.
3.5.2.2 Laser Spot Jitter
Laser spot jitter is defined as the high frequency motion of the laser spot on a pulse-to-pulse basis, usually
of low amplitude, and ostensibly due to minute flexures of the optical bench caused by aircraft vibration.
These rapid angular movements of the beam degrade weapon accuracy only slightly when the laser beam
is normal to the target face. However, at shallow grazing angles and large slant-ranges, jitter may cause
each spot to move hundreds of feet in relation to the aim point and in relation to the previous spot location.
In many cases (e.g., most self-designation LGB deliveries), this movement is near perpendicular to the
weapon flight path and create false left-right commands. Therefore, as the weapon manoeuvres to intercept
the moving spot, this factor may cause rapid depletion of the LGB available energy and may cause large
miss distances to be generated.
3.5.2.3 Laser Boresight Error
Laser boresight error is defined as the misalignment between the location of the aiming reticle and the
laser spot on the target. This error is easy to visualize as a geometric progression of the beam wandering
away from the sensor sight line as the range increases. Boresight error is not only a static error source but

LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 25


can be a dynamic error as well. The system optical bench may distort, changing the designator/sensor
boresight relation as the system is slewed through its field of regard. In addition, manoeuvring (g forces)
may cause additional shifts as the structure between the designator and sensor deflects under load. In some
cases, particularly at long slant-ranges, boresight error can place the laser spot off the target, resulting in a
weapon miss. If the magnitude of boresight error is known, however, the aimpoint can be shifted to
compensate.
3.5.2.4 Laser Pointing Error
Laser pointing error is defined as the inability to place the laser spot at the exact desired location on the
target. This is usually observed when trying to designate a small target from long ranges, where the reticle
size can obscure the target. If the sensor magnification of the target is insufficient, it is difficult to know
exactly where the aiming reticle is located on the target and, sometimes, it may be also difficult to know if
it is on the target at all.
3.5.2.5 Tracking Error
Tracking error is a generic term that encompasses other forms of spot movement from the desired aim point.
Where jitter is a random movement of the beam around a central axis, tracking error may be described as
undesired movement of this central axis around or away from the aim point. This movement of the central
beam axis may or may not be visible to the operator depending on the magnitude of the error and the quality
of the sensor presentation to the operator. At long slant-ranges, automatic tracking systems can exhibit beam
wander that overwhelms other sources of error. This wander is caused by movement of the video tracking
gates on-or-about the aimpoint as the viewing aspect changes. The changing aspect or look angle produces
changes in the aim point contrast with respect to its background. This, in turn, varies the location of the
contrast driven tracking gate position with a consequent shift in beam position. Other causes for tracking
error may include g forces (mentioned earlier), transient angle rate errors due to rapid bank angle changes,
or momentary errors due to LOS masking. Motion of the laser spot during the last three seconds prior to
impact may induce unnecessary corrections to the weapon flight and result in a miss.
3.5.3 Podium Effect
For an LGB to guide, the seeker must be in a position to receive the reflected laser energy. During a self-
designation attack against a vertical target, there is a risk that the laser spot will move around the target
face relative to the weapon LOS, as the designator aircraft flies the recovery manoeuvre, and that the
weapon will not receive the reflected laser energy during the final critical moments before impact. This
phenomena, known as the podium effect, is particularly apparent when the designator to target line is
significantly different to that of the weapons flight path. To avoid the podium effect, the designating
aircraft should maneuver such that the target face is always in front of the aircraft and that the appropriate
terminal slant-range/angle occurs at weapon impact. This problem can often be eliminated by lasing on top
of a horizontal target.
3.5.4 Beam Divergence and Reflected Power
Another effect of beam divergence is to reduce the maximum reflected power available to the weapon as
the beam strikes the target off-axis. Figure 3-8 illustrates the laser spot shape and intensity versus various
designation angles of incidence. The calculations assume a 100% diffuse surface, no atmospheric
attenuation, and an illuminating beam with a Gaussian distribution.
LASER SYSTEMS PERFORMANCE
3 - 26 RTO-AG-300-V26























Figure 3-8: Laser Spot Intensity vs. Angle of Incidence.
3.5.5 Sensor Resolution
The size of the target must also be factored against the resolution abilities of the sensor element (FLIR
and/or TV) to determine the maximum usable delivery slant-range. This will ensure that the operator will
be able to resolve the target at a range that is in excess of the maximum range capability of the weapon.
This excess or redundant range requirement is necessary to properly detect and then identify the target
prior to weapon release. This target detection and identification requirement prior to release has become of
almost paramount importance in punitive or other high visibility actions where the blind launches required
by other weapon systems prevent their use.
As mentioned earlier, the maximum slant-range from which a designator is intended to be operated must
be determined as part of the mission planning process as a function of target size, laser system error
budget, and laser power. In addition, an attempt should be made to determine what additional range should
be selected in order for the target to be properly identified prior to weapon release. This requires an
estimate of the time required to first detect the target on the sensor set and then add the time required to
100% Peak Intensity
86.6% Peak Intensity
50% Peak Intensity
34% Peak Intensity
x
2x
3x
y
y
1.16x
y
y
0 Incidence
30 Incidence
60 Incidence
70 Incidence
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 27


fully resolve the target for a positive identification. With current TV/FLIR technologies and good initial
cueing, it is usually estimated that at least ten seconds are required to detect the target. Further five to ten
seconds are then required to properly identify the target itself.
3.5.6 Airborne LTD/LGB Mission Geometry
Let us consider again the LTD/LGW attack geometry already described in Figure 3-1. With reference to
this geometry, the maximum range performance of an LTD/LGB combination can be estimated using the
Eq. (3.21), which we write again:

2 2 2
4
R T T L
atm R r t T
R ) R D (
cos cos cos UA
MDED


+
= (3.71)
Conveniently, in Eq. (3.71), we have replaced the term
| | ) (
R HR T HR w
R R
e
+
(i.e., two-ways atmospheric
transmittance) with the symbol
atm
, and the returned energy density (I ) with the Minimum Detectable
Energy Density (MDED) of the LGB seeker-head unit.
There are three cosine factors in Eq. (3.71). They are related to the assumption of a Lambertian reflection
(i.e., diffuse reflection of the laser signal incident on the target surface). It is important, in order to
determine the performance of an LTD/LGW combination during an attack, to take into account the
variations of the angles
t
,
r
and
r
. On the other hand, in order to calculate the maximum range for an
effective illumination in the worst geometric case, it is important to determine the maximum values
assumed by these angles during the attack. Moreover, for mission planning purposes, it is useful to express
the angles
t
,
r
and
r
as functions of other physical or geometrical parameters that are known prior the
mission (e.g., seeker FOV, target inclination). Using Eq. (3.76), the maximum theoretical value of the
angle
r
can be determined as a function of the seeker Minimum Detectable Energy Density (MDED).
However, we must consider that the seeker of the LGW must always intercept a portion of the reflected
signal sufficient to produce a response of the detector in order to guide the weapon against the target.
In other words, the angle
r
(MDED) should always be greater than the FOV of the seeker (see Figure 3-9).

Figure 3-9: LGB-Target Geometry.

r
(MDED)
FOV

r

LASER SYSTEMS PERFORMANCE
3 - 28 RTO-AG-300-V26


Considering the geometry of typical ground attack missions with LGB, the angles
t
(angle between the
LOS transmitter-target and the normal to the target surface) and
r
(angle between the LOS receiver-target
and the normal to the target surface), can be expressed as function of other geometric parameters and their
maximum theoretical values (corresponding to the minimum relative range performance) can be
determined. With reference to Figure 3-2, the angles
t
and
r
can be expressed as:

2

+ =
t t
i (3.72)

r r
i

=
2
(3.73)
where i is the target inclination,
t
is the angle between the transmitted beam axis and the horizon and
r
is
the angle between the LGW-target LOS and the horizon (
r t t r
= ). Knowing
d
, and , it is
possible to determine the value of the angle
t
during the attack, solving the equation:

2

+ + =
d t
i (3.74)
More difficult is the determination of
r
, since the angle
r
can not be determined without knowing
continuously the position assumed by the line of sight LGW-target (i.e., the guidance algorithms and
corrected ballistics of the LGW). However, knowing the angle at the beginning of the designation (from
the ballistics of the unguided weapon) and taking
r
equivalent to the seeker FOV, we have that:
FOV
) MAX ( r r
= = (3.75)
Since it is reasonable to assume that, after the designation is initiated, the angle
r
will be kept as low as
possible by a PG-LGW, we can assume that
r
in this case.
Therefore, the approximate value of the angle
r
during an attack with PG-LGB and BTB-LGB, can be
determined solving the equations:
= i
r
90 for PG-LGW (3.76)
FOV i
r
+ = 90 for BTB-LGW (3.77)
For the purpose of determining the maximum values that the angles
t
and
r
can reach during an attack,
which determine the absolute minimum performance of a particular LTD/LGB combination (worst case),
it is meaningful to take into account the tactics of typical self-designation attacks illustrated in Figure 3-10.
Since the designation is initiated in the final portion of the bomb trajectory (i.e., with an LTD-target range
typically between 1.2 and 2.0 times the release range), it is generally performed at a considerable range from
the target. This means that, normally, the angles
t
and
r
never reach values close to 90 during an attack,
even in the worst case when i = 90. On the other hand, in the case of horizontal target (i = 0), the cases
where
t
and
r
are close to 90 are of little practical interest.


LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 29



Figure 3-10: LTD/LGB Mission Horizontal Profiles (Self-Designation).
Looking at Figure 3-11, it appears evident that the angle
t
is smaller than i when i > 45, while it is
generally smaller than the complementary of i when i < 45. Similar considerations apply to
r
. Therefore,
with these assumptions, the worst case conditions for
t
and
r
are the following:

=
=
i
i
) MAX ( r
) MAX ( t
2
2

for i < 45 ;

=
=
i
i
) MAX ( r
) MAX ( t

for i 45 (3.78)

Target
TACTIC N 2
TACTIC N 1
LTD Position at
Weapon Impact
Weapon
Release
Target
Target
Maximum Designation
Range (2R
R
)
Target
LASER SYSTEMS PERFORMANCE
3 - 30 RTO-AG-300-V26




45
45
i
i = 0
90-i
90-i
= t limit
= r limit
i
90
90
i = 60
i = 30
i = 45
i = 90

Figure 3-11: Limits of the Angles
t
and
r
.
3.5.7 LTD System Error Budget
As an example, we consider a LGB which can achieve a 10 feet Circular Error Probability (CEP). In this
case, it is appropriate that the designator aiming capability must equal or exceed that requirement in order
to meet a suitable weapon impact criteria. If a hardened shelter access cover, roughly 20 feet in diameter,
is considered as a target in our example, this target dimension equates to a 10 feet CEP where 50 % of our
hypothetical weapon releases should fall on the target face. Using Tactic 2 shown in Figure 3-10 against a
vertical target, and choosing a desired release range (R
R
) of 35,000 feet, it is necessary that our designator
must be capable of keeping its beam on a 20 feet diameter target at a Terminal Slant-Range (TSR) of
70,000 feet. This equates to a total allowable Maximum Error Budget (EB
max
) of 285 rad (20 ft / 70 Kft).
We also assume that the target is designated at the corresponding terminal designation angle () of 60
off of the line normal to the target face. This 60 offset reduces the gross error budget to approximately
143 rad (EB
max
cos60). This means that all pointing and beam divergence error sources, when added in
a worst case fashion, must fall within a cone that subtends 143 rad if 50% of our hypothetical weapons
are to hit the 20 feet target mentioned above.
In the light of the above considerations, the maximum allowable error budget can be expressed as:

TSR
cos T
EB
S
max

= (3.79)
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 31


where T
S
is the target size and TSR is the Terminal Slant-Range. Using Tactic 2 in Figure 3-10, the terminal
slant-range can be expressed as:

cos
R
TSR
R
= (3.80)
3.5.8 Release Range
Given a fixed error budget and known designation tactic (e.g., Tactic 2), we can solve for the optimal release
range:

( )
2
max
S
R
EB
cos T
R

= (3.81)
Using for example a worst case error budget of 208 rad (given by the sum of all pointing error
contributions), the optimal release range against a 20 feet target with a 60 terminal designation angle is
approximately 24 Kft (i.e., not 35 Kft as originally desired). This example demonstrates that, in most cases
with LGW, the engagement scenario is usually limited by designator and/or sensor capability, and not by
the standoff capability of the weapon itself, particularly at extreme slant-ranges and/or low graze angles.
3.5.9 Maximum Egress Range
Due to the tracking error of the LTD system described above, a 600 kts ingress would require approximately
15 to 20 thousand feet of additional range over that of the desired release range. In other words, a 600 KTAS
ingress to a 35,000 foot release point would require a detection range of over 50,000 to 55,000 feet.
However, both designation and sensor capabilities should be geared toward the egress side of the picture.
During egress, the designator aircraft would desirably turn to a heading that provides maximum standoff
and yet provide a flight path that will stay within designator constraints up until weapon impact. With
reference to Figure 3-10 (showing two possible tactics that might be used), Tactic 1 is probably the most
desirable in terms of standoff, however, it requires a designator with full hemispheric coverage below the
aircraft for high altitude delivery or full coverage above the aircraft for low altitude deliveries. Tactic 2
shows a probable tactic that could be used when a rear gimbal limit has been placed on the LTD aiming
system. While standoff is probably acceptable, a major constraint then becomes the look angle at a vertical
target face from the LTD perspective (Podium Effect). As the designator proceeds outbound after weapon
release, the perceived horizontal dimension of the target decreases by up to 50 percent (for an optimum
attack heading). Where the attack heading is constrained and an optimum attack solution is not available,
the off axis perspective may reduce one target dimension by another 20%.
Ordinarily, as in both of the above cases, the range attained during egress is normally greater than the
ingress range required for detection. For present LGB weapons, the range during egress at weapon impact
time typically varies from approximately 1.2 to 2.0 times the release range. This ratio shifts towards 2.0 as
standoff is increased towards maximum range. For the example given earlier, the designator aircraft would
be at a slant-range of between 42,000 and 70,000 feet at weapon impact.
3.5.10 Masking
Another important problem with airborne laser systems is masking of the equipment field of regard caused
by the aircraft structure and loads (e.g., weapons, external tanks). Although masking can be reduced/
eliminated by a careful aircraft/system design in the case of embedded systems, this is generally a very
important constraint for operations with podded systems (e.g., the CLDP integrated on the Italian
TORNADO-IDS). A useful way of characterising systems masking characteristics is the so called Masking
LASER SYSTEMS PERFORMANCE
3 - 32 RTO-AG-300-V26


Matrix. This is a Cartesian co-ordinate system in which (most conveniently) azimuth and elevation are
plotted for the equivalent FOV of the system. This is given by intersection of the system visibility matrix
and the aircraft matrix (e.g., an aircraft/loads CAD model). For the airborne LTD system in service with
the Italian Air Force (CLDP), the system masking is essentially given by a backward cone with an aperture
of 30 and 20, for the IR and TV front sections respectively (Figure 3-12).



Figure 3-12: CLDP FOV Limitations (TV and IR).
During the CLDP integration on TORNADO-IDS, analysis was required in order to fully characterise the
masking phenomenon and obtain the related mathematical model to be used by the aircraft MC for CLDP
inhibition during impingement.
The initial TORNADO-IDS masking model (developed by ALENIA) was obtained through a computer
CAD simulation, that consisted in defining the aircraft shape with different external stores configurations.
As a result of the analysis/simulation, the proposed masking function logic was defined (Figure 3-13).
IR

TV
20
30
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 33














ROOT
LO
RO
LSW
TANK
LI
RI
E/N
CL
E/N
A/C MASKS
SEMI
CLEAN 2
TANK
E/N
GBU24
GBU24
E/N
E/N
E/N
CLEAN
SYMBOLS:
E/N = Empty / No Pylon
TANK = 1500 LT Tank
GBU24 = GBU24 Real Weapon
CBLS = BDU 33 B/B or MK 106
= Dont Care
TANK
RS (light)
CBLS
GBU16
RS (heavy) E/N
GBU16
SEMI
CLEAN 1
E/N
RSW



Figure 3-13: CLDP Masking Selection Logic.
Particularly, the aircraft masking function was conceived in order to manage the basic real GBU-16 and
GBU-24 Stores Configurations (worst case masking profile) and their derived sub-configurations
(i.e., semi-clean and clean), providing appropriate aural/visual warning to the pilot/WSO and inhibition
commands to the CLDP laser in case of LOS impingement with aircraft and stores. Furthermore,
a pre-masking function was implemented in order to provide aural/visual advice to the pilot/WSO in
case of approximation to the masking conditions.
The validity of the solutions developed for masking/attack profiles and Laser illumination phase, was
verified through simulation and flight tests (jointly by ALENIA and the Italian Air Force). The developed
simulation tool, fitted with the suitable problem oriented routines, allowed the exploration of the system
behaviour under the influence of a large number of parameters. Particularly, simulation was used to
monitoring the LOS components (Azimuth and Elevation) in an Hammer/Aitoff diagram where mask and
pre-mask conditions were plotted.
The trajectory of the LOS, marked in time between the bomb release and impact, gave an immediate
understanding about the effect of the aircraft manoeuvre on the LOS pointing direction. By varying the
aircraft manoeuvre parameters (i.e., turning direction, turning load factor, roll rate, and egress heading),
the LOS trajectory gave an indication on the critical conditions that could arise with the chosen
parameters.
LASER SYSTEMS PERFORMANCE
3 - 34 RTO-AG-300-V26


The basic software tool was composed by an aircraft mathematical dynamics model, based on the classical
equations set, used in conjunction with a simplified aircraft data bank containing the main TORNADO-
IDS characteristics. The aircraft model was provided with a simplified autopilot able to maintain flight
path parameters (i.e., height, velocity and heading) aimed at performing automatic attack manoeuvres
(e.g., turns, climbs, dives), used during the evaluation phase.
Furthermore, a simplified program that simulated the LGB ballistic trajectory was used. This code run as a
stand alone task and was used to compute in advance range and time of flight of the bomb for the chosen
release conditions. These data were then loaded into the simulator memory to command the post weapon
delivery manoeuvre.
Flight test activities performed by ALENIA and the Italian Air Force Official Flight Test Centre (RSV),
permitted to finally tune and validate the masking and pre-masking algorithms [15]. Particularly, tests
were conducted in selected portions of the operational flight envelopes, representative of real LTD/LGB
attack missions and of the boundary conditions for activation of the masking and pre-masking functions.
3.6 REFERENCES
[1] Jelalian, A.V., Laser Radar Systems. Artech House Boston-London. 1992.
[2] Sabatini, R., Tactical Laser Systems Performance Prediction in Various Weather Conditions.
1
st
Symposium of the NATO-RTO SET Panel (former AGARD-SPP Panel). Italian Air Force
Academy. Naples (Italy). 16-19 March 1998.
[3] Weichel, H., Laser Beam Propagation in the Atmosphere. SPIE Optical Engineering Press. Second
Printing. 1990.
[4] Elder, T. and Strong, J., The Infrared Transmission of Atmospheric Windows. J. Franklin Institute
255-189. 1953.
[5] Langer, R.M., Signal Corps Report N DA-36-039-SC-72351. May 1957.
[6] Kneizys, F.X., Shuttle, E.P., Abreau, L.W., Chetwynd, J.H., Anderson, G.P., Gallery, W.O.,
Selby, J.E.A. and Clough, S.A., Users Guide to LOWTRAN 7. Air Force Geophysical Laboratory
Report AFGL-TR-88-0177. Hansom AFB (MA). 1988.
[7] Middleton, W.E.K., Vision Through the Atmosphere. University of Toronto Press. 1952.
[8] Middleton, W.E.K., Vision Through the Atmosphere. Handbuch der Physik 48. Geophysics 2.
Springer (Berlin). 1957.
[9] Hudson, R.D., Infrared Systems Engineering. Wiley & Sons. 1969.
[10] Holst, G.C., Electro-Optical Imaging System Performance. SPIE Optical Engineering Press.
Bellingham, Washington USA. 1995.
[11] Chu, T.S. and Hogg, D.C., Effects of Precipitation on Propagation at 0.63, 3.5 and 10.6 Microns.
Bell Systems Technical Journal 47 No. 5. 1968.
[12] Strohbehn, J.W. et al., Laser Beam Propagation in the Atmosphere. Topics in Applied Physics
Series Vol. 25. Springer-Verlag. 1978.
LASER SYSTEMS PERFORMANCE
RTO-AG-300-V26 3 - 35


[13] Keith, G.G., Otten, L.J. and Rose, W.C., Aerodynamic Effects. ERIM-SPIE IR&EO Systems
Handbook (Vol. 2 Chapter 3). Second Printing. 1996.
[14] Phong, B.T., Illumination for Computer Generated Pictures. Communications of the ACM. Vol.
18-6 (pp. 311-317). 1975.
[15] Sabatini, R., Guercio, F., Marciante, A. and Campo, G., Laser Guided Bombs and Convertible
Designation Pod Integration on Italian TORNADO-IDS. 31
st
Annual Symposium of the Society of
Flight Test Engineers. Turin (Italy). 18-22 September 2000.
LASER SYSTEMS PERFORMANCE
3 - 36 RTO-AG-300-V26

You might also like